• Aucun résultat trouvé

Accelerating MCMC algorithms

N/A
N/A
Protected

Academic year: 2021

Partager "Accelerating MCMC algorithms"

Copied!
22
0
0

Texte intégral

(1)

Accelerating MCMC Algorithms

Christian P Robert

Universit´e Paris Dauphine, PSL Research University, and Department of Statistics, University of Warwick

V´ıctor Elvira

IMT Lille Douai & CRIStAL laboratory

Nick Tawn

Department of Statistics, University of Warwick

Changye Wu

Universit´e Paris Dauphine, PSL Research University

Abstract. Markov chain Monte Carlo algorithms are used to simulate from complex statistical distributions by way of a local exploration of these distributions. This local feature avoids heavy requests on un-derstanding the nature of the target, but it also potentially induces a lengthy exploration of this target, with a requirement on the number of simulations that grows with the dimension of the problem and with the complexity of the data behind it. Several techniques are available towards accelerating the convergence of these Monte Carlo algorithms, either at the exploration level (as in tempering, Hamiltonian Monte Carlo and partly deterministic methods) or at the exploitation level (with Rao-Blackwellisation and scalable methods).

1. INTRODUCTION

Markov chain Monte Carlo (MCMC) algorithms have been used for nearly 60 years, becoming a reference method for analysing Bayesian complex models in the early 1990’s (Gelfand and Smith, 1990). The strength of this method is that it guarantees convergence to the quantity (or quantities) of interest with minimal requirements on the targeted distribution (also called target) behind such quantities. In that sense, MCMC algorithms are robust or universal, as opposed to the most standard Monte Carlo methods (see, e.g., Rubinstein, 1981; Robert and Casella, 2004) that require direct simulations from the target distribution. This robustness may however induce a slow convergence behaviour in that the exploration of the relevant space—meaning the part of the space supporting the distribution that has a significant probability mass under that distribution— may take a long while, as the simulation usually proceeds by local jumps in the vicinity of the current position. In other words, MCMC–especially in its off-the-shelf versions like Gibbs sampling and Metropolis–Hastings algorithms—is very often myopic in that it provides a good illumination of a local area, while remaining unaware of the global support of the distribution. As with most other

1

(2)

simulation methods, there always exist ways of creating highly convergent MCMC algorithms by taking further advantage of the structure of the target distribution. Here, we mostly limit ourselves to the realistic situation where the target density is only known as the output of a computer code or to a setting similarly limited in its information content.

The approaches to the acceleration of MCMC algorithms can be divided in several categories, from those which improve our knowledge about the target distribution, to those that modify the proposal in the algorithm, including those that exploit better the outcome of the original MCMC algorithm. The following sections provide more details about these directions and the solutions proposed in the literature.

2. WHAT IS MCMC AND WHY DOES IT NEED ACCELERATING? MCMC methods have a history (see, e.g.Capp´e and Robert,2000) that starts at approximately the same time as the Monte Carlo methods, in conjunction with the conception of the first computers. They have been devised to handle the simulation of complex target distributions, when complexity stems from the shape of the target density, the size of the associated data, the dimension of the object to be simulated, or from time requirements. For instance, the target density π(θ) may happen to be expressed in terms of multiple integrals that cannot be solved analytically,

π(θ) = Z

ω(θ, ξ)dξ ,

which requires the simulation of the entire vector (θ, ξ). In cases when ξ is of the same dimension as the data, as for instance in latent variable models, this significant increase in the dimension of the object to be simulated creates compu-tational difficulties for standard Monte Carlo methods, from managing the new target ω(θ, ξ), to devising a new and efficient simulation algorithm. A Markov chain Monte Carlo (MCMC) algorithm allows for an alternative resolution of this computational challenge by simulating a Markov chain that explores the space of interest (and possibly supplementary spaces of auxiliary variables) with-out requiring a deep preliminary knowledge on the density π, besides the ability to compute π(θ0) for a given parameter value θ0 (if up to a normalising constant)

and possibly the gradient ∇ log π(θ0). The validation of the method (e.g.,Robert

and Casella,2004) is that the Markov chain is ergodic (e.g., Meyn and Tweedie, 1993), namely that it converges in distribution to the distribution with density π, no matter where the Markov chain is started at time t = 0.

The Metropolis–Hastings algorithm is a generic illustration of this principle. The basic algorithm is constructed by choosing a proposal, that is, a conditional density K(θ′|θ) (also known as a Markov kernel), the Markov chain {θ

t}∞t=1being

then derived by successive simulations of the transition

θt+1=    θ′ ∼ K(θ t) with probability  π(θ ′ ) π(θt) ×K(θt|θ ′ ) K(θ′ t)  ∧ 1, θt otherwise.

This acceptance-rejection feature of the algorithm makes it appropriate for tar-geting π as its stationary distribution if the resulting Markov chain {θt}∞t=1 is

(3)

irreducible, i.e., has a positive probability of visiting any region of the support of π in a finite number of iterations. (Stationarity can easily be shown, e.g., by using the so-called detailed balance property that makes the chain time-reversible, see, e.g.,Robert and Casella,2004.)

Considering the initial goal of simulating samples from the target distribution π, the performances of MCMC methods like the Metropolis–Hastings algorithm above often vary quite a lot, depending primarily on the correspondance between the proposal K and the target π. For instance, if K(θ|θt) = π(θ), the Metropolis–

Hastings algorithm reduces to i.i.d. sampling from the target, which is of course a formal option when i.i.d. sampling from π proves impossible to implement. Al-though there exist rare instances when the Markov chain {θt}∞t=1leads to negative

correlations between the successive terms of the chain, making it more efficient than regular i.i.d. sampling (Liu et al., 1995), the most common occurrence is one of positive correlation between the simulated values (sometimes uniformly, see Liu et al., 1994). This feature implies a reduced efficiency of the algorithm and hence requires a larger number of simulations to achieve the same precision as an approximation based on i.i.d. simulations (without accounting for differ-ences in computing time). More generally, a MCMC algorithm may require a large number of iterations to escape the attraction of its starting point θ0 and

to reach stationarity, to the extent that some versions of such algorithms fail to converge in the time available (i.e., in practice if not in theory).

It thus makes sense to seek ways of accelerating (a) the convergence of a given MCMC algorithm to its stationary distribution, (b) the convergence of a given MCMC estimate to its expectation, and/or (c) the exploration of a given MCMC algorithm of the support of the target distribution. Those goals are related but still distinct. For instance, a chain initialised by simulating from the target dis-tribution may still fail to explore the whole support in an acceptable number of iterations. While there is not an optimal and universal solution to this issue, we will discuss below approaches that are as generic as possible, as opposed to artificial ones taking advantage of the mathematical structure of a specific tar-get distribution. Ideally, we aim at covering realistic situations when the tartar-get density is only known [up to a constant or an additional completion step] as the output of an existing computer code. Pragmatically, we also cover here solutions that require more efforts and calibration steps when they apply to a wide enough class of problems.

3. ACCELERATING MCMC BY EXPLOITING THE GEOMETRY OF THE TARGET

While there is no end in trying to construct more efficient and faster MCMC algorithms, and while this (endless) goal needs to account for the cost of devising such alternatives under limited resources budgets, there exist several generic so-lutions such that a given target can first be explored in terms of the geometry (or topology) of the density before constructing the algorithm. Although this type of methods somehow takes us away from our original purpose which was to improve upon an existing algorithm, they still make sense within this survey in that they allow for almost automated implementations.

(4)

3.1 Hamiltonian Monte Carlo

From the point of view of this review, Hamiltonian (or hybrid) Monte Carlo (HMC) is an auxiliary variable technique that takes advantage of a continuous time Markov process to sample from the target π. This approach comes from physics (Duane et al., 1987) and was popularised in statistics by Neal (1999, 2011) andMacKay(2002). Given a target π(θ), where θ ∈ Rd, an artificial

auxil-iary variable ϑ ∈ Rd is introduced along with a density ̟(ϑ|θ) so that the joint distribution of (θ, ϑ) enjoys π(θ) as its marginal. While there is complete free-dom in this representation, the HMC literature often calls ϑ the momentum of a particle located at θ by analogy with physics. Based on the representation of the joint distribution

ω(θ, ϑ) = π(θ)̟(ϑ|θ) ∝ exp{−H(θ, ϑ)} ,

where H(·) is called the Hamiltonian, Hamiltonian Monte Carlo (HMC) is asso-ciated with the continuous time process (θt, ϑt) generated by the so-called

Hamil-tonian equations dθt dt = ∂H ∂ϑ(θt, ϑt) dϑt dt = − ∂H ∂θ (θt, ϑt) , which keep the Hamiltonian target stable over time, as

dH(θt, ϑt) dt = ∂H ∂ϑ(θt, ϑt) dϑt dt + ∂H ∂θ (θt, ϑt) dθt dt = 0 .

Obviously, the above continuous time Markov process is deterministic and only explores a given level set,

{(θ, ϑ) : H(θ, ϑ) = H(θ0, ϑ0)} ,

instead of the whole augmented state space R2d, which induces an issue with irreducibility. An acceptable solution to this problem is to refresh the momentum, ϑt ∼ ̟(ϑ|θt−), at random times {τn}∞n=1, where θt− denotes the location of θ

immediately prior to time t, and the random durations {τn− τn−1}∞n=2 follow an

exponential distribution. By construction, continuous-time Hamiltonian Markov chain can be regarded as a specific piecewise deterministic Markov process using Hamiltonian dynamics (Davis,1984,1993;Bou-Rabee et al.,2017) and our target, π, is the marginal of its associated invariant distribution.

Before moving to the practical implementation of the concept, let us point out that the free cog in the machinery is the conditional density ̟(ϑ|θ), which is usually chosen as a Gaussian density with either a constant covariance matrix M corresponding to the target covariance or as a local curvature depending on θ in Riemannian Hamiltonian Monte Carlo (Girolami and Calderhead,2011). Betan-court(2017) argues in favour of these two cases against non-Gaussian alternatives and Livingstone et al. (2017) analyse how different choices of kinetic energy in Hamiltonian Monte Carlo affect algorithm performances. For a fixed covariance matrix, the Hamiltonian equations become

dθt

dt = M

−1ϑ

t dϑt

(5)

which is the score function. The velocity (or momentum) of the process is thus driven by this score function, gradient of the log-target.

The above description remains quite conceptual in that there is no generic methodology for producing this continuous time process, since the Hamiltonian equations cannot be solved exactly in most cases. Furthermore, standard numer-ical solvers like Euler’s method create an instable approximation that induces a bias as the process drifts away from its true trajectory. There exists however a discretisation simulation technique that produces a Markov chain and is well-suited to the Hamiltonian equations in that it preserves the stationary distribu-tion (Betancourt,2017). It is called the symplectic integrator, and one version in the independent case with constant covariance consists in the following (so-called leapfrog) steps

ϑt+ǫ/2 = ϑt+ ǫ∇L(θt)/2,

θt+ǫ= θt+ ǫM−1ϑt+ǫ/2,

ϑt+ǫ= ϑt+ǫ/2+ ǫ∇L(θt+ǫ)/2,

where ǫ is the time-discretisation step. Using a proposal on ϑ0 drawn from the

Gaussian auxiliary target and deciding on the acceptance of the value of (θT ǫ, ϑT ǫ)

by a Metropolis–Hastings step can limit the danger of missing the target. Note that the first two leapfrog steps induce a Langevin move on θt:

θt+ǫ= θt+ ǫ2M−1∇L(θt)/2 + ǫM−1ϑt,

thus connecting with the MALA algorithm discussed below (see Durmus and Moulines,2017for a theoretical discussion of the optimal choice of ǫ). Note that the leapfrog integrator is quite an appealing middleground between accuracy (as it is second-order accurate) and computational efficiency.

In practice, it is important to note that discretising the Hamiltonian dynamics introduces two free parameters, the step size ǫ and the trajectory length T ǫ, both to be calibrated. As an empirically successful and popular variant of HMC, the “no-U-turn sampler” (NUTS) ofHoffman and Gelman(2014) adapts the value of ǫ based on primal-dual averaging. It also eliminates the need to choose the trajec-tory length T via a recursive algorithm that builds a set of candidate proposals for a number of forward and backward leapfrog steps and stops automatically when the simulated path steps back.

A further acceleration step in this area is proposed byRasmussen (2003) (see also Fielding et al., 2011), namely the replacement of the exact target density π(·) by an approximation ˆπ(·) that is much faster to compute in the many iter-ations of the HMC algorithm. A generic way of constructing this approximation is to rely on Gaussian processes, when interpreted as prior distributions on the target density π(·), which is only observed at some values of θ, π(θ1), . . . , π(θn)

(Rasmussen and Williams, 2005). This solution is speeding up the algorithm, possibly by orders of magnitude, but it introduces a further approximation into the Monte Carlo approach, even when the true target is used at the end of the leapfrog discretisation, as inFielding et al.(2011).

Stan (named after Stanislas Ullam, see Carpenter et al.,2017) is a computer language for Bayesian inference that, among other approximate techniques, im-plements the NUTS algorithm to remove hand-tuning. More precisely, Stan is a

(6)

probabilistic programming language in that the input is at the level of a statis-tical model, along with data, rather than the specifics of an MCMC algorithm. The algorithmic part is somehow automated, meaning that when models can be conveniently defined through this language, it offers an alternative to the sampler that produced the original chain. As an illustration of the acceleration brought by HMC, Figure1, reproduced fromHoffman and Gelman(2014), shows the per-formance of NUTS, compared with both random-walk MH and Gibbs samplers.

Fig 1: Comparisons between random-walk Metropolis-Hastings, Gibbs sampling, and NUTS algorithm of samples corresponding to a highly correlated 250-dimensional multivariate Gaussian target. Similar computation budgets are used for all methods to produce the 1,000 samples on display (Source: Hoffman and Gelman(2014), with permission).

4. ACCELERATING MCMC BY BREAKING THE PROBLEM INTO PIECES

The explosion in the collection and analysis of “big” datasets in recent years has brought new challenges to the MCMC algorithms that are used for Bayesian inference. When examining whether or not a new proposed sample is accepted at the accept-reject step, an MCMC algorithm such as the Metropolis-Hastings ver-sion needs to sweep over the whole data set, at each and every iteration, for the evaluation of the likelihood function. MCMC algorithms are then difficult to scale up, which strongly hinders their application in big data settings. In some cases, the datasets may be too large to fit on a single machine. It may also be that con-fidentiality measures impose different databases to stand on separate networks, with the possible added burden of encrypted data (Aslett et al.,2015). Commu-nication between the separate machines may prove impossible on an MCMC scale that involves thousands or hundreds of thousands iterations.

4.1 Scalable MCMC methods

In the recent years, efforts have been made to design scalable algorithms, namely, solutions that manage to handle large scale targets by breaking the problem into manageable or scalable pieces. Roughly speaking, these methods can be classified into two categories (Bardenet et al., 2015): divide-and-conquer approaches and sub-sampling approaches.

Divide-and-conquer approaches partition the whole data set, denoted X , into batches, {X1, · · · , Xk}, and run separate MCMC algorithms on each data batch,

(7)

independently, as if they were independent Bayesian inference problems.1 These methods then combine the simulated parameter outcomes together to approx-imate the original posterior distribution. Depending on the treatments of the batches selected in the MCMC stages, these approaches can be further subdivided into two finer groups: sub-posterior methods and boosted sub-posterior methods. Sub-posterior methods are motivated by the independent product equation:

(1) π(θ) ∝ k Y i=1  π0(θ)1/k Y ℓ∈Xi p(xℓ|θ)  = k Y i=1 πi(θ) ,

and they target the densities πi(θ) (up to a constant) in their respective MCMC

steps. They thus bypass communication costs (Scott et al., 2016), by running MCMC samplers independently on each batch, and they most often increase MCMC mixing rates (in effective samples sizes produced by second), given that the sub-posterior distributions πi(θ) are based on smaller datasets. For instance,

Scott et al. (2016) combine the samples from the sub-posteriors, πi(θ), by a

Gaussian reweighting.Neiswanger et al. (2013) estimate the sub-posteriors πi(θ)

by non-parametric and semi-parametric methods, and they run additional MCMC samplers on the product of these estimators towards approximating the true posterior π(θ). Wang and Dunson (2013) refine this product estimator with an additional Weierstrass sampler, while Wang et al. (2015) estimate the posterior by partitioning the space of samples with step functions. Vehtari et al. (2014) devised an expectation propagation scheme to improve the postprocessing of the parallel samplers.

As an alternative to sampling from the sub-posteriors, boosted sub-posterior methods target instead the components

(2) π˜i(θ) ∝ π0(θ)   Y ℓ∈Xi p(xℓ|θ)   k

in separate MCMC runs. Since they formaly amount to repeating each batch k times towards producing pseudo data sets with the same size as the true one, the resulting boosted sub-posteriors, ˜π1(θ), · · · , ˜πk(θ), have the same scale in

vari-ance of each component of the parameters, θ, as the true posterior, and can thus be treated as a group of estimators of the true posterior. In the subsequent combining stage, these sub-posteriors are merged together to construct a better approximation of the target distribution. For instance, Minsker et al.(2014) ap-proximate the posterior with the geometric median of the boosted sub-posteriors, embedding them into associated reproducing kernel Hilbert spaces (rkhs), while Srivastava et al.(2015) achieve this goal using the barycentres of ˜π1, · · · , ˜πk, these

barycentres being computed with respect to a Wasserstein distance.

In a perspective different from the above parallel scheme of divide-and-conquer approaches, sub-sampling approaches aim at reducing the number of individual datapoint likelihood evaluations operated at each iteration towards accelerating MCMC algorithms. From a general perspective, these approaches can be further

1In order to keep the notations consistent, we still denote the target density by π, with the

prior density denoted as π0 and the sampling distribution of one observation x as p(x|θ). The

(8)

classified into two finer classes: exact subsampling methods and approximate subsampling methods, depending on their resulting outputs. Exact subsampling approaches typically require subsets of data of random size at each iteration. One solution to this effect is taking advantage of pseudo-marginal MCMC via constructing unbiased estimators of the target density evaluated on subsets of the data (Andrieu and Roberts,2009). Quiroz et al.(2016) follow this direction by combining the powerful debiasing technique of Rhee and Glynn (2015) and the correlated pseudo-marginal MCMC approach of Deligiannidis et al. (2015). Another direction is to use piecewise deterministic Markov processes (PDMP) (Davis,1984,1993), which enjoy the target distribution as the marginal of their invariant distribution. This PDMP version requires unbiased estimators of the gradients of the log-likelihood function, instead of the likelihood itself. By us-ing a tight enough bound on the event rate function of the associated Poisson processes, PDMP can produce super-efficient scalable MCMC algorithms. The bouncy particle sampler (Bouchard-Cˆot´e et al., 2017) and the zig-zag sampler (Bierkens et al.,2016) are two competing PDMP algorithms, whileBierkens et al. (2017) unify and extend these two methods. Besides, one should note that PDMP produces a non-reversible Markov chain, which means that the algorithm should be more efficient in terms of mixing rate and asymptotic variance, when compared with reversible MCMC algorithms, such as MH, HMC and MALA, as observed in some theoretical and experimental works (Hwang et al.,1993;Sun et al.,2010; Chen and Hwang,2013;Bierkens,2016).

Approximate subsampling approaches aim at constructing an approximation of the target distribution. Beside the aforementioned attempts of Rasmussen (2003) andFielding et al.(2011), one direction is to approximate the acceptance probability with high accuracy by using subsets of the data (Bardenet et al., 2014,2015). Another solution is based on a direct modification of exact methods. The seminal work of Welling and Teh(2011), SGLD, is to exploit the Langevin diffusion (3) dθt= 1 2ΛΛΛ∇ log π(θt)dt + ΛΛΛ 1/2dB t, θ0∈ Rd, t ∈ [0, ∞)

where ΛΛΛ is a user-specified matrix, π is the target distribution and Bt is a

d-dimensional Brownian process. By virtue of the Euler-Maruyama discretisation and using unbiased estimators of the gradient of the log-target density, SGLD and its variants (Ding et al., 2014; Chen et al., 2014) often produce fast and accurate results in practice when compared with MCMC algorithms using MH steps.

Figure 2 shows the time requirements of a consensus Monte Carlo algorithm (Scott et al., 2016) compared with a Metropolis–Hastings algorithm using the whole dataset, while Figure 3 displays the saving in likelihood evaluations in confidence sampler of Bardenet et al.(2015).

4.2 Parallelisation and distributed schemes

Modern computational architectures are built with several computing units that allow for parallel processing, either fully independent or with certain com-munication. Although the Markovian nature of MCMC is inherently sequential and somewhat alien to the notion of parallelising, several partial solutions have been proposed in the literature for exploiting these parallel architectures. The

(9)

Fig 2: Elapsed time when drawing 10,000 MCMC samples with different amounts of data under the single machine and consensus Monte Carlo algorithms for a hierarchical Poisson regression. The horizontal axis represents the amounts of data. The single machine algorithm stops after 30 because of the explosion in computation budget. (Source:Scott et al.(2016), with permission.)

simplest approach consists in running several MCMC chains in parallel, blind to all others, until the allotted computing time is exhausted. Finally, the resulting estimators of all chains are averaged. However, this naive implementation may suffer from the fact that some of those chains have not reached their station-ary regime by the end of the computation time, which then induces a bias in the resulting estimate. Ensuring that stationarity has been achieved is a difficult (if at all possible) task, although several approaches can be found in the litera-ture (Mykland et al.,1995;Guihenneuc-Jouyaux and Robert,1998;Jacob et al., 2017). At the opposite extreme, complex targets may be represented as products that involve many terms that must be evaluated, each of which can be attributed to a different thread before being multiplied all together. This strategy requires communication among processors at each MCMC step. A middle-ground version (Jacob et al., 2011) consists in running several Markov chains in parallel with periodic choices of the reference chain, all simulations being recycled through a Rao-Blackwell scheme. (See also Calderhead,2014 for a similar scheme.) The family of interacting orthogonal MCMC methods (O-MCMC) is proposed in Mar-tino et al.(2016) with the aim of fostering better exploration of the state space, specially in high-dimensional and multimodal targets. Multiple MCMC chains are run in parallel exploring the space with random-walk proposals. The paral-lel chains periodically share information, also through joint MCMC steps, thus allowing an efficient combination of global (coordinated) exploration and local ap-proximation. O-MCMC methods also allow for a parallel implementation of the Multiple Try Metropolis (MTM). In Calderhead (2014), a generalisation of the Metropolis-Hastings algorithm allows for a straightforward parallelisation. Each proposed point can be evaluated in a different processor at every MCMC itera-tion. Finally, note that the section on scalable MCMC also contains parallelisable

(10)

Fig 3: Percentage of numbers of data points used in each iteration of the confi-dence sampler with a single 2nd order Taylor approximation at θMAP. The plots

describe 10,000 iterations of the confidence sampler for the posterior distribu-tion of the mean and variance of a uni-dimensional Normal distribudistribu-tion with a flat prior: (left) 10,000 observations are generated from N (0, 1), (right) 10,000 observations are generated from LN (0, 1) (Source: Bardenet et al. (2015), with permission).

approaches, such as the prefetching method of Angelino et al. (2014) (see also Banterle et al.,2015for a related approach, primarily based on an approximation of the target). A most recent endeavour called asynchronous MCMC (Terenin et al., 2015) aims at higher gains in parallelisation by reducing the amount of exchange between the parallel threads, but the notion still remains confidential at this stage.

5. ACCELERATING MCMC BY IMPROVING THE PROPOSAL In the same spirit as the previous section, this section is stretching the pur-pose of this paper by considering possible modifications of the MCMC algorithm itself, rather than merely exploiting the output of a given MCMC algorithm. For instance, devising an HMC algorithm is an answer to this question even though the “improvement” is not garanteed. Nonetheless, our argument here is that, once provided with this output, it is possible to derive new proposals in a semi-autonomous manner.

5.1 Simulated tempering

The target distribution, π(θ) on d-dimensional state space Θ, can exhibit multi-modality with the probability mass being located in different regions in the state space. The majority of MCMC algorithms use a localised proposal mechanism which is tuned towards local approximate optimality see, e.g., Roberts et al. (1997) andRoberts and Rosenthal (2001). By construction, these localised pro-posals result in the Markov chain becoming “trapped” in a subset of the state space meaning that in finite run-time the chain can entirely fail to explore other modes in the state space, leading to biased samples. Strategies to accelerate MCMC often use local gradient information and this draws the chain back to-wards the centre of the mode, which is the opposite of what is required in a multi-modal setting.

There is an array of methodology available to overcome issues of multi-modality in MCMC, the majority of which use state space augmentation. Auxiliary

(11)

dis-tributions that allow a Markov chain to explore the entirety of the state space are targeted and their mixing information is then passed on to aid mixing in the true target. While the sub-posteriors of the previous section can be seen as special cases of the following, the most successful and convenient implementation of these methods is to use power-tempered target distributions. The target distribution at inverse temperature level, β, for β ∈ (0, 1] is defined as

πβ(θ) = K(β) [π(θ)]β where K(β) =

Z

[π(θ)]βdθ −1

.

Therefore, π1(θ) = π(θ). Temperatures β < 1 flatten out the target distribution

allowing the chain to explore the entire state space provided the β value is suffi-ciently small. The simulated tempering (ST) and parallel tempering (PT) algo-rithms (Geyer,1991;Marinari and Parisi,1992) typically use the power-tempered targets to overcome the issue of multi-modality. The ST approach runs a single Markov chain on the augmented state space {B, Θ}, where B = {β0, β1, . . . , βn}

is a discrete collection of n inverse temperature levels with 1 = β0 > β1 > . . . >

βn > 0. The algorithm uses a Metropolis-within-Gibbs strategy by cycling

be-tween updates in the Θ and B components of the space. For instance, a proposed temperature swap move βi → βj is accepted with probability

min 

1,πβj(θ)

πβi(θ)



in order to preserve detailed balance. Note that this acceptance ratio depends on the normalisation constants K(β) which are typically unknown, although they can sometimes be estimated, as in, e.g.,Wang and Landau (2001) and Atchad´e and Liu (2004). In case estimation of the marginal normalisation constants is impractical then the PT algorithm is employed. This approach simultaneously runs a Markov chain at each of the n + 1 temperature levels targeting the joint distribution given byQn

i=0[π(θi)]βi. Swap moves between chains at adjacent

tem-perature levels are accepted according to a ratio that no longer depends on the marginal normalisation constants. Indeed, this power tempering approach has been successfully employed in a number of settings and is widely used e.g.Neal (1996), Earl and Deem (2005), Xie et al. (2010), Mohamed et al. (2012) and Carter and White(2013).

In both approaches, there is a “Goldilocks” principle to setting up the inverse temperature schedule. Spacings between temperature levels that are “too large” result in swap moves that are rarely accepted, hence delaying the transfer of hot state mixing information to the cold states. On the other hand, spacings that are too small require a large number of intermediate temperature levels, again resulting in slow mixing through the temperature space. This problem becomes even more difficult as the dimensionality of Θ increases.

Much of the historical literature suggested that a geometric spacing was opti-mal i.e., there exists c ∈ (0, 1) such that βi+1 = cβi for i = 0, 1, . . . , n. However,

in the case of the simulated tempering version (ST),Atchad´e et al.(2011) consid-ered the problem as an optimal scaling problem by maximising the (asymptotic in dimension) expected squared jumping distance in the B space for temperature swap moves. Under restrictive assumptions, he showed that the spacings between

(12)

consecutive inverse temperature levels should scale with dimension as O d−1/2 to prevent degeneracy of the swap move acceptance rate. For a practitioner the result gave guidance on optimal setup since it suggested a corresponding opti-mal swap move acceptance rate of 0.234 between consecutive inverse temperature levels, in accordance withGelman et al. (1996). Finally, contrary to the histori-cally recommended geometric schedule, the authors suggested that temperature schedule setup should be constructed consecutively so as to induce an approxi-mate 0.234 swap acceptance rate between consecutive levels; which is achieved adaptively in Miasojedow et al. (2013). The use of expected squared jumping distance as the measure of mixing speed was justified in Roberts and Rosenthal (2014) where, under the same conditions as inAtchad´e et al.(2011), it was shown that the temperature component of the ST chain has an associated diffusion pro-cess.

The target of an 0.234 acceptance rate gives good guidance to setting up the ST/PT algorithms in certain settings, but there is a major warning for practition-ers following this rule for optimal setup. The assumptions made inAtchad´e et al. (2011) andRoberts and Rosenthal(2014) ignore the restrictions of mixing within a temperature level, instead assuming that this can be done infinitely fast relative to the mixing within the temperature space. Woodard et al. (2009a), Woodard et al. (2009b) and Bhatnagar and Randall (2016) undertake a comprehensive analysis of the spectral gap of the ST/PT chains and their conclusion is rather damning of the ST/PT approaches that use power-tempered targets. Essentially, in situations where the modes have different structures, the time required to reach a given level of convergence for the ST/PT algorithms can grow exponen-tially in dimension. A major reason for this is that power-based tempering does not preserve the relative weights/mass between regions at the different temper-ature levels, see Figure4. This issue can scale exponentially in dimension. From a practical perspective, in these finite run high-dimensional non-identical modal structure settings the swap acceptance rates can be very misleading, meaning that they have limited use as a diagnostic for inter-modal mixing quality. 5.2 Adaptive MCMC

Improving and calibrating an MCMC algorithm towards a better correspon-dance with the intended target is a natural step in making the algorithm more efficient, provided enough information is available about this target distribution. For instance, when an MCMC sample associated with this target is available, even when it has not fully explored the range of the target, it contains some amount of information, which can then be exploited to construct new MCMC algorithms. Some of the solutions available in the literature (see, e.g.Liang et al., 2007) proceed by repeating blocks of MCMC iterations and updating the pro-posal K after each block, aiming at a particular optimality goal like a specific acceptance rate like 0.234 for Metropolis–Hastings steps (Gelman et al., 1996). Most versions of this method update the scale structure of a random walk pro-posal, based on previous realisations (Robert and Casella,2009) or on an entire sample (Douc et al., 2007a), which turns the method into iterated importance sampling with Markovian dependence. (It can also be seen as a static version of particle filtering,Doucet et al.,2000;Andrieu and Doucet,2002;Storvik,2002.) Other adaptive resolutions bypass this preliminary and somewhat ad hoc

(13)

con-Fig 4: Un-normalised tempered target densities of a bimodal Gaussian mixture using inverse temperature levels β = {1, 0.1, 0.05, 0.005} respectively. At the hot state (bottom right) it is evident that the mode centred on 40 begins to domi-nate the weight as β increases to ∞ even though at the cold state it was only attributable for a fraction (0.2) of the total mass.

struction and aim instead at a permanent updating within the algorithm, moti-vated by the idea that a continuous adaptation keeps improving the correspon-dance with the target. In order to preserve the validation of the method (Gelman et al., 1996;Haario et al., 1999; Roberts and Rosenthal,2007;Saksman and Vi-hola,2010), namely that the chain produced by the algorithm converges to the intended target, specific convergence results need be established, as the ergodic theorem behind standard MCMC algorithms do not apply. Without due caution (see Figure 5), an adaptive MCMC algorithm may fail to converge due to over-fitting. A drawback of adaptivity is that the update of the proposal distribution relies too much on the earlier simulations and thus reinforces the exclusion of parts of the space that have not yet been explored.

For the validation of adaptive MCMC methods, stricter constraints must thus be imposed on the algorithm. One well-described solution (Roberts and Rosen-thal,2009) is called diminishing adaptation. Informally, it consists in imposing a distance between two consecutive proposal kernels to uniformly decrease to zero. In practice, this means stabilising the changes in the proposal by ridge-like factors as in the early proposal byHaario et al. (1999). A drawback of this resolution is that the decrease itself must be calibrated and may well fail to bring a significant improvement over the original proposal.

5.3 Multiple try MCMC

A completely different approach to improve the original proposal used in an MCMC algorithm is to consider a collection of proposals, built on different ra-tionales and experiments. The multiple try MCMC algorithm (Liu et al., 2000; B´edard et al.,2012;Martino,2018) follows this perspective. As the name suggests, the starting point of a multiple try MCMC algorithm is to simultaneously pro-pose N potential moves θ1

t, . . . , θNt of the Markov chain, instead of a single value.

(14)

differ-Fig 5: Markov chains produced by an adaptive algorithm where the proposal distribution is a Gaussian distribution with mean and variance computed from the past simulations of the chain. The three rows correspond to different initial distributions. The fit of the histogram of the resulting MCMC sample is poor, even for the most spread-out initial distribution (bottom) (Robert and Casella, 2004, with permission)

(15)

ent proposal densities Ki(·|θt) that are conditional on the current value of the

Markov chain, θt. One of the θti’s is selected based on the importance sampling

weights wi

t ∝ π(θti)/Ki(·|θt). The selected value is then accepted by a further

Metropolis–Hastings step which involves a ratio of normalisation constants for the importance stage, one corresponding to the selection made previously and another one created for this purpose. Indeed, besides the added cost of com-puting the sum of the importance weights and generating the different variates, this method faces the non-negligible drawback of requiring N − 1 supplementary simulations that are only used for achieving detailed balance and computing a backward summation of importance weights. This constraint may vanish when considering a collection of independent Metropolis-Hastings proposals, q(θ), but this setting is rarely realistic as it requires which make life simpler, but are less realistic since some amount of prior knowledge or experimentation to build a relevant distribution.

An alternative found in the literature is ensemble Monte Carlo (Iba, 2000; Capp´e et al., 2008; Neal, 2011; Martino, 2018), illustrated in Figure 6 which produces a whole sample at each iteration, with target the product of the initial targets, in closer proximity with particle methods (Capp´e et al.,2004;Mengersen and Robert,2003).

Fig 6: A comparison of an ensemble MCMC approach with a regular adaptive MCMC algorithm (lower line) and a static importance sampling approach, in terms of mean square error (MSE), for a fixed total number of likelihood eval-uations, where N denotes the size of the ensemble (Source: Martino, 2018, with permission).

Yet another implementation of this principle is called delayed rejection ( Tier-ney and Mira, 1998; Mira, 2001; Mira and Sargent, 2003), where proposals are instead considered sequentially, once the previous proposed value has been re-jected. to speeding up MCMC by considering several possibilities, if sequentially. A computational difficulty with this approach is that the associated acceptance probabilities get increasingly complex as the number of delays grows, which may annihilate its appeal relative to simultaneous multiple tries. A further difficulty

(16)

is to devise the sequence of proposals in a diverse enough manner. 5.4 Multiple proposals and parameterisations

A rather basic approach to comparing proposals of MCMC algorithms is to run several in parallel and to check whether these parallel chains can be exchanged by coupling. Chains with divergent behaviour will not couple as often as chains exploring the same area. While creating multiple MCMC algorithms may seem a major challenge, automated and semi-automated schemes can be replicated as much as desired by changing the parameterisation of the target. Each change introduces a different Jacobian in the expression of the density, which means different efficiencies in the exploration of the target.

6. ACCELERATING MCMC BY REDUCING THE VARIANCE

Since the main goal of MCMC is to produce approximations for quantities of interest of the form

Ih= Z

Θ

h(θ)π(θ)dθ,

an alternative (and cumulative) way of accelerating these algorithms is to improve the quality of the approximation derived from an MCMC output. That is, given an MCMC sequence θ1, . . . , θT, converging to π(·), one can go beyond resorting

to the basic Monte Carlo approximation

(4) ˆIT h =1/T T X t=1 h(θt)

towards reducing the variance (if not the speed of convergence) of ˆIT

h to Ih.

A common remark when considering Monte Carlo approximations of Ihis that

the representation of the integral as an expectation is not unique (e.g. Robert and Casella, 2004). This leads to the technique of importance sampling where alternative distributions are used in replacement of π(θ), possibly in an adaptive manner (Douc et al.,2007b), or sequentially as in particle filters (Del Moral et al., 2006; Andrieu et al., 2011). Within the framework of this essay, the outcome of a given MCMC sampler can also be exploited in several ways that lead to an improvement of the approximation of Ih.

6.1 Rao–Blackwellisation and other averaging techniques

The name ‘Rao–Blackwellisation’ was coined byGelfand and Smith (1990) in their foundational Gibbs sampling paper and it has since then become a stan-dard way of reducing the variance of integral approximations. While it essentially proceeds from the basic probability identity

Eπ[h(θ)] = Eπ1[Eπ2{h(θ)|ξ}],

when π can be expressed as the following marginal density π(θ) =

Z

Ξ

π1(ξ)π2(θ|ξ)dξ ,

and while sufficiency does not have a clear equivalence for Monte Carlo approxi-mation, the name stems from the Rao–Blackwell theorem (Lehmann and Casella,

(17)

1998) that improves upon a given estimator by conditioning upon a sufficient statistics. In a Monte Carlo setting, this means that (4) can be improved by a partly integrated version

(5) I˜Th =1/T

T

X

t=1

Eπ2[h(θ)|ξt]

assuming that a second and connected sequence of simulations (ξt) is available

and that the conditional expectation is easily constructed. For instance, Gibbs sampling (Gelfand and Smith,1990) is often open to this Rao–Blackwell decompo-sition as it relies on successive simulations from several conditional distributions, possibly including auxiliary variates and nuisance parameters. In particular, a generic form of Gibbs sampling called the slice sampler (Robert and Casella, 2004) produces one or several uniform variates at each iteration.

However, a more universal type of Rao–Blackwellisation is available (Casella and Robert,1996) for all MCMC methods involving rejection, first and foremost, Metropolis–Hastings algorithms. Indeed, first, the distribution of the rejected variables can be derived or approximated, which leads to an importance correc-tion of the original estimator. Furthermore, the accept-reject step depends on a uniform variate, but this uniform variate can be integrated out. Namely, given a sample produced by a Metropolis–Hastings algorithm θ(1), . . . , θ(T ), one can

ex-ploit both underlying samples, the proposed values ϑ1, . . . , ϑT, and the uniform

u1, . . . , uT, so that the ergodic mean can be rewritten as

ˆ IT h =1/T T X t=1 h(θ(t)) =1/T T X t=1 h(ϑt) T X i=t Iθ(i) t.

The conditional expectation ˜ IT h =1/T T X t=1 h(ϑt)E " T X i=t Iθ(i) t ϑ1, . . . , ϑT # =1/T T X t=1 h(ϑt) ( T X i=t P(θ(i)= ϑt|ϑ1, . . . , ϑT) )

then enjoys a smaller variance. See alsoTjelmeland(2004) andDouc and Robert (2010) for connected improvements based on multiple tries. An even more rudi-mentary (and cheaper) version can be considered by integrating out the decision step at each Metropolis–Hastings iteration: if θtis the current value of the Markov

chain and ϑt the proposed value, to be accepted (as θt+1) with probability αt,

the version 1/T T X t=1 {αth(ϑt) + (1 − αt)h(θt)}

should most often2 bring an improvement over the basic estimate (Liu et al.,

1995;Robert and Casella,2004).

2The improvement is not universal, due to the correlation between the terms of the sum

(18)

7. CONCLUSION

Accelerating MCMC algorithms may sound like a new Achille versus tortoise paradox in that there are aways methods to speed up a given algorithm. The stopping rule of this infinite regress is however that the added pain in achieving this acceleration may overcome the added gain at some point. While we have only and mostly superficially covered some of the possible directions in this survey, we thus encourage most warmly readers to keep an awareness for the potential brought by a wide array of almost cost-free accelerating solutions as well as to keep trying devising more fine-tuned improvements in every new MCMC imple-mentation. For instance, for at least one of us, Rao-Blackwellisation is always considered at this stage. Keeping at least one such bag of tricks at one’s disposal is thus strongly advised.

ACKNOWLEDGEMENTS

Christian P. Robert is grateful to Gareth Roberts, Mike Betancourt, and Julien Stoehr for helpful discussions. He is currently supported by an Institut Universi-taire de France 2016–2021 senior grant. Changye Wu is currently a Ph.D. candi-date at Universit´e Paris-Dauphine and supported by a grant of the Chinese Gov-ernment (CSC). V´ıctor Elvira acknowledges support from the Agence Nationale de la Recherche of France under PISCES project (ANR-17-CE40-0031-01), the Fulbright program, and the Marie Curie Fellowship (FP7/2007-2013) under REA grant agreement n. PCOFUND-GA-2013-609102, through the PRESTIGE pro-gram. The authors are quite grateful to a reviewer for his or her detailed coverage of an earlier version of the paper, which contributed to significant improvements in the presentation and coverage of the topic. All remaining errors and ommissions are solely the responsability of the authors.

REFERENCES

Andrieu, C.and Doucet, A. (2002). Particle filtering for partially observed Gaussian state space models. J. Royal Statist. Society Series B, 64 827–836.

Andrieu, C., Doucet, A. and Holenstein, R. (2011). Particle Markov chain Monte Carlo (with discussion). J. Royal Statist. Society Series B, 72 (2) 269–342.

Andrieu, C.and Roberts, G. (2009). The pseudo-marginal approach for efficient Monte Carlo computations. The Annals of Statistics 697–725.

Angelino, E., Kohler, E., Waterland, A., Seltzer, M. and Adams, R. (2014). Accelerating MCMC via parallel predictive prefetching. arXiv preprint arXiv:1403.7265.

Aslett, L., Esperanc¸a, P. and Holmes, C. (2015). A review of homomorphic encryption and software tools for encrypted statistical machine learning. arXiv:1508.06574.

Atchad´e, Y. F.and Liu, J. S. (2004). The Wang-Landau algorithm for Monte Carlo compu-tation in general state spaces. Statistica Sinica, 20 209–33.

Atchad´e, Y. F., Roberts, G. and Rosenthal, J. (2011). Towards optimal scaling of Metropolis-coupled Markov chain Monte Carlo. Statistics and Computing, 21 555–568. Banterle, M., Grazian, C., Lee, A. and Robert, C. P. (2015). Accelerating Metropolis–

Hastings algorithms by delayed acceptance. arXiv preprint arXiv:1503.00996.

Bardenet, R., Doucet, A. and Holmes, C. (2014). Towards scaling up Markov chain Monte Carlo: an adaptive subsampling approach 405–413.

Bardenet, R., Doucet, A. and Holmes, C. (2015). On Markov chain Monte Carlo methods for tall data. J. Machine Learning Res., 18 1515–1557.

B´edard, M., Douc, R. and Moulines, E. (2012). Scaling analysis of multiple-try MCMC methods. Stochastic Processes and their Applications, 122 758–786.

Betancourt, M. (2017). A conceptual introduction to Hamiltonian Monte Carlo. ArXiv

(19)

Bhatnagar, N.and Randall, D. (2016). Simulated tempering and swapping on mean-field models. Journal of Statistical Physics, 164 495–530.

Bierkens, J.(2016). Non-reversible Metropolis–Hastings. Statistics and Computing, 26 1213– 1228.

Bierkens, J., Bouchard-Cˆot´e, A., Doucet, A., Duncan, A. B., Fearnhead, P., Roberts, G.and Vollmer, S. J. (2017). Piecewise deterministic Markov processes for scalable Monte Carlo on restricted domains. arXiv preprint arXiv:1701.04244.

Bierkens, J., Fearnhead, P. and Roberts, G. (2016). The zig-zag process and super-efficient sampling for Bayesian analysis of big data. arXiv preprint arXiv:1607.03188.

Bou-Rabee, N., Sanz-Serna, J. M. et al. (2017). Randomized hamiltonian monte carlo.

The Annals of Applied Probability, 27 2159–2194.

Bouchard-Cˆot´e, A., Vollmer, S. J. and Doucet, A. (2017). The bouncy particle sampler: A non-reversible rejection-free Markov chain Monte Carlo method. Journal of the American

Statistical Association.

Calderhead, B. (2014). A general construction for parallelizing Metropolis–Hastings algo-rithms. Proceedings of the National Academy of Sciences, 111 17408–17413.

Capp´e, O., Douc, R., Guillin, A., Marin, J.-M. and Robert, C. (2008). Adaptive impor-tance sampling in general mixture classes. Statist. Comput., 18 447–459.

Capp´e, O., Guillin, A., Marin, J.-M. and Robert, C. (2004). Population Monte Carlo. J.

Comput. Graph. Statist., 13 907–929.

Capp´e, O.and Robert, C. (2000). Ten years and still running! J. American Statist. Assoc., 95 1282–1286.

Carpenter, B., Gelman, A., Hoffman, M., Lee, D., Goodrich, B., Betancourt, M., Brubaker, M., Guo, J., Li, P. and Riddell, A. (2017). Stan: A probabilistic programming language. Journal of Statistical Software, Articles, 76.

Carter, J.and White, D. (2013). History matching on the imperial college fault model using parallel tempering. Computational Geosciences, 17 43–65.

Casella, G.and Robert, C. (1996). Rao-Blackwellization of sampling schemes. Biometrika, 83 81–94.

Chen, T., Fox, E. and Guestrin, C. (2014). Stochastic gradient hamiltonian monte carlo 1683–1691.

Chen, T.and Hwang, C. (2013). Accelerating reversible Markov chains. Statistics & Probability

Letters, 83 1956–1962.

Davis, M. H.(1984). Piecewise-deterministic Markov processes: A general class of non-diffusion stochastic models. Journal of the Royal Statistical Society. Series B (Methodological) 353–388. Davis, M. H.(1993). Markov Models & Optimization, vol. 49. CRC Press.

Del Moral, P., Doucet, A. and Jasra, A. (2006). Sequential Monte Carlo samplers. J.

Royal Statist. Society Series B, 68 411–436.

Deligiannidis, G., Doucet, A. and Pitt, M. K. (2015). The correlated pseudo-marginal method. arXiv preprint arXiv:1511.04992.

Ding, N., Fang, Y., Babbush, R., Chen, C., Skeel, R. D. and Neven, H. (2014). Bayesian sampling using stochastic gradient thermostats 3203–3211.

Douc, R., Guillin, A., Marin, J.-M. and Robert, C. (2007a). Convergence of adaptive mixtures of importance sampling schemes. Ann. Statist., 35(1) 420–448.

Douc, R., Guillin, A., Marin, J.-M. and Robert, C. (2007b). Convergence of adaptive mixtures of importance sampling schemes. Ann. Statist., 35(1) 420–448. ArXiv:0708.0711. Douc, R. and Robert, C. (2010). A vanilla variance importance sampling via population

Monte Carlo. Ann. Statist. To appear.

Doucet, A., Godsill, S. and Andrieu, C. (2000). On sequential Monte-Carlo sampling methods for Bayesian filtering. Statist. Comp., 10 197–208.

Duane, S., Kennedy, A. D., Pendleton, B. J. and Roweth, D. (1987). Hybrid Monte Carlo. Phys. Lett. B, 195 216–222.

Durmus, A.and Moulines, E. (2017). Nonasymptotic convergence analysis for the unadjusted Langevin algorithm. Ann. Applied Probability, 27 1551–1587.

Earl, D. J.and Deem, M. W. (2005). Parallel tempering: Theory, Applications, and New Perspectives. Physical Chemistry Chemical Physics, 7 3910–3916.

Fielding, M., Nott, D. J. and Liong, S.-Y. (2011). Efficient MCMC schemes for computa-tionally expensive posterior distributions. Technometrics, 53 16–28. https://doi.org/10. 1198/TECH.2010.09195, URLhttps://doi.org/10.1198/TECH.2010.09195.

(20)

Gelfand, A.and Smith, A. (1990). Sampling based approaches to calculating marginal den-sities. J. American Statist. Assoc., 85 398–409.

Gelman, A., Gilks, W. and Roberts, G. (1996). Efficient Metropolis jumping rules. In

Bayesian Statistics 5 (J. Berger, J. Bernardo, A. Dawid, D. Lindley and A. Smith, eds.).

Oxford University Press, Oxford, 599–608.

Geyer, C. J.(1991). Markov chain Monte Carlo maximum likelihood. Computing Science and

Statistics, 23 156–163.

Girolami, M.and Calderhead, B. (2011). Riemann manifold Langevin and Hamiltonian Monte Carlo methods. Journal of the Royal Statistical Society: Series B (Statistical

Method-ology), 73 123–214.

Guihenneuc-Jouyaux, C. and Robert, C. P. (1998). Discretization of continuous Markov chains and Markov chain Monte Carlo convergence assessment. Journal of the American

Statistical Association, 93 1055–1067.

Haario, H., Saksman, E. and Tamminen, J. (1999). Adaptive proposal distribution for random walk Metropolis algorithm. Computational Statistics, 14(3) 375–395.

Hoffman, M. D.and Gelman, A. (2014). The No-U-turn sampler: adaptively setting path lengths in Hamiltonian Monte Carlo. J. Machine Learning Res., 15 1593–1623.

Hwang, C.-R., Hwang-Ma, S.-Y. and Sheu, S.-J. (1993). Accelerating gaussian diffusions.

The Annals of Applied Probability 897–913.

Iba, Y. (2000). Population-based Monte Carlo algorithms. Trans. Japanese Soc. Artificial

Intell., 16 279–286.

Jacob, P. E., O’Leary, J. and Atchad´e, Y. F.(2017). Unbiased Markov chain Monte Carlo with couplings. ArXiv e-prints. 1708.03625.

Jacob, P., Robert, C. P. and Smith, M. H. (2011). Using parallel computation to improve independent Metropolis–Hastings based estimation. Journal of Computational and Graphical

Statistics, 20 616–635.

Lehmann, E.and Casella, G. (1998). Theory of Point Estimation (revised edition). Springer-Verlag, New York.

Liang, F., Liu, C. and Carroll, R. (2007). Stochastic approximation in Monte Carlo com-putation. JASA, 102 305–320.

Liu, J., Wong, W. and Kong, A. (1994). Covariance structure of the Gibbs sampler with application to the comparison of estimators and augmentation schemes. Biometrika, 81 27– 40.

Liu, J., Wong, W. and Kong, A. (1995). Covariance structure and convergence rates of the Gibbs sampler with various scans. Journal of Royal Statistical Society, B 57 157–169. Liu, J. S., Liang, F. and Wong, W. H. (2000). The multiple-try method and local optimization

in Metropolis sampling. J. American Statist. Assoc., 95 121–134.

Livingstone, S., Faulkner, M. F. and Roberts, G. O. (2017). Kinetic energy choice in hamiltonian/hybrid monte carlo. arXiv preprint arXiv:1706.02649.

MacKay, D. J. C.(2002). Information Theory, Inference & Learning Algorithms. Cambridge University Press, Cambridge, UK.

Marinari, E.and Parisi, G. (1992). Simulated tempering: a new Monte Carlo scheme. EPL

(Europhysics Letters), 19 451.

Martino, L.(2018). A Review of Multiple Try MCMC algorithms for Signal Processing. ArXiv

e-prints. 1801.09065.

Martino, L., Elvira, V., Luengo, D., Corander, J. and Louzada, F. (2016). Orthogonal parallel MCMC methods for sampling and optimization. Digital Signal Processing, 58 64–84. Mengersen, K. and Robert, C. (2003). Iid sampling with self-avoiding particle filters: the pinball sampler. In Bayesian Statistics 7 (J. Bernardo, M. Bayarri, J. Berger, A. Dawid, D. Heckerman, A. Smith and M. West, eds.). Oxford University Press, Oxford.

Meyn, S.and Tweedie, R. (1993). Markov Chains and Stochastic Stability. Springer-Verlag, New York.

Miasojedow, B., Moulines, E. and Vihola, M. (2013). An adaptive parallel tempering algorithm. Journal of Computational and Graphical Statistics, 22 649–664.

Minsker, S., Srivastava, S., Lin, L. and Dunson, D. (2014). Scalable and robust Bayesian inference via the median posterior 1656–1664.

Mira, A.(2001). ”on metropolis-hastings algorithms with delayed rejection”. Metron, 59 (3-4) 231–241.

(21)

algorithms. Stat. Methods Appl., 12 49–60.

Mohamed, L., Calderhead, B., Filippone, M., Christie, M. and Girolami, M. (2012). Population mcmc methods for history matching and uncertainty quantification.

Computa-tional Geosciences, 16 423–436.

Mykland, P., Tierney, L. and Yu, B. (1995). Regeneration in Markov chain samplers. Journal

of the American Statistical Association, 90 233–241.

Neal, R.(1999). Bayesian Learning for Neural Networks, vol. 118. Springer–Verlag, New York. Lecture Notes.

Neal, R.(2011). MCMC using Hamiltonian dynamics. In In Handbook of Markov Chain Monte

Carlo (S. Brooks, A. Gelman, G. L. Jones and X.-L. Meng, eds.). CRC Press, New York.

Neal, R. M. (1996). Sampling from Multimodal Distributions using Tempered Transitions.

Statistics and Computing, 6 353–366.

Neal, R. M. (2011). MCMC Using Ensembles of States for Problems with Fast and Slow Variables such as Gaussian Process Regression. ArXiv e-prints. 1101.0387.

Neiswanger, W., Wang, C. and Xing, E. (2013). Asymptotically exact, embarrassingly par-allel MCMC. arXiv preprint arXiv:1311.4780.

Quiroz, M., Villani, M. and Kohn, R. (2016). Exact subsampling MCMC. arXiv preprint

arXiv:1603.08232.

Rasmussen, C. E.(2003). Gaussian processes to speed up hybrid monte carlo for expensive bayesian integrals. In Bayesian Statistics 7 (J. Bernardo, M. Bayarri, J. Berger, A. Dawid, D. Heckerman, A. Smith and M. West, eds.). Oxford University Press, 651–659.

Rasmussen, C. E.and Williams, C. K. I. (2005). Gaussian Processes for Machine Learning

(Adaptive Computation and Machine Learning). The MIT Press.

Rhee, C.-h.and Glynn, P. W. (2015). Unbiased estimation with square root convergence for sde models. Operations Research, 63 1026–1043.

Robert, C.and Casella, G. (2004). Monte Carlo Statistical Methods. 2nd ed. Springer-Verlag, New York.

Robert, C. and Casella, G. (2009). Introducing Monte Carlo Methods with R. Springer-Verlag, New York.

Roberts, G., Gelman, A. and Gilks, W. R. (1997). Weak convergence and optimal scaling of random walk Metropolis algorithms. The Annals of Applied Probability, 7 110–120. Roberts, G. and Rosenthal, J. (2001). Optimal scaling for various Metropolis-Hastings

algorithms. Statistical Science, 16 351–367.

Roberts, G.and Rosenthal, J. (2007). Coupling and ergodicity of adaptive Markov Chain Monte carlo algorithms. J. Applied Proba., 44(2) 458–475.

Roberts, G.and Rosenthal, J. (2009). Examples of adaptive MCMC. J. Comp. Graph. Stat., 18 349–367.

Roberts, G.and Rosenthal, J. (2014). Minimising MCMC variance via diffusion limits, with an application to simulated tempering. The Annals of Applied Probability, 24 131–149. Rubinstein, R. Y.(1981). Simulation and the Monte Carlo Method. J. Wiley, New York. Saksman, E.and Vihola, M. (2010). On the ergodicity of the adaptive Metropolis algorithm

on unbounded domains. Ann. Applied Probability, 20(6) 2178–2203.

Scott, S. L., Blocker, A. W., Bonassi, F. V., Chipman, H. A., George, E. I. and McCul-loch, R. E.(2016). Bayes and big data: The consensus Monte Carlo algorithm. International

Journal of Management Science and Engineering Management, 11 78–88.

Srivastava, S., Cevher, V., Dinh, Q. and Dunson, D. (2015). Wasp: Scalable Bayes via barycenters of subset posteriors 912–920.

Storvik, G.(2002). Particle filters for state space models with the presence of static parameters.

IEEE Trans. Signal Process., 50 281–289.

Sun, Y., Schmidhuber, J. and Gomez, F. J. (2010). Improving the asymptotic performance of Markov chain Monte-carlo by inserting vortices 2235–2243.

Terenin, A., Simpson, D. and Draper, D. (2015). Asynchronous Gibbs Sampling. ArXiv

e-prints. 1509.08999.

Tierney, L.and Mira, A. (1998). Some adaptive Monte Carlo methods for Bayesian inference.

Statistics in Medicine, 18 2507–2515.

Tjelmeland, H. (2004). Using all Metropolis–Hastings proposals to estimate mean values. Tech. Rep. 4, Norwegian University of Science and Technology, Trondheim, Norway. Vehtari, A., Gelman, A., Sivula, T., Jyl¨anki, P., Tran, D., Sahai, S., Blomstedt, P.,

(22)

as a way of life: A framework for Bayesian inference on partitioned data. ArXiv e-prints.

1412.4869.

Wang, F.and Landau, D. (2001). Determining the density of states for classical statistical models: A random walk algorithm to produce a flat histogram. Physical Review E, 64 056101. Wang, X.and Dunson, D. (2013). Parallelizing MCMC via weierstrass sampler. arXiv preprint

arXiv:1312.4605.

Wang, X., Guo, F., Heller, K. and Dunson, D. (2015). Parallelizing MCMC with random partition trees. Advances in Neural Information Processing Systems 451–459.

Welling, M.and Teh, Y. (2011). Bayesian learning via stochastic gradient Langevin dynamics 681–688.

Woodard, D. B., Schmidler, S. C. and Huber, M. (2009a). Conditions for rapid mixing of parallel and simulated tempering on multimodal distributions. The Annals of Applied

Probability 617–640.

Woodard, D. B., Schmidler, S. C. and Huber, M. (2009b). Sufficient conditions for torpid mixing of parallel and simulated tempering. Electronic Journal of Probability, 14 780–804. Xie, Y., Zhou, J. and Jiang, S. (2010). Parallel tempering monte carlo simulations of lysozyme

orientation on charged surfaces. The Journal of chemical physics, 132 02B602.

Figure

Fig 1: Comparisons between random-walk Metropolis-Hastings, Gibbs sampling, and NUTS algorithm of samples corresponding to a highly correlated  250-dimensional multivariate Gaussian target
Fig 2: Elapsed time when drawing 10,000 MCMC samples with different amounts of data under the single machine and consensus Monte Carlo algorithms for a hierarchical Poisson regression
Fig 3: Percentage of numbers of data points used in each iteration of the confi- confi-dence sampler with a single 2nd order Taylor approximation at θ MAP
Fig 4: Un-normalised tempered target densities of a bimodal Gaussian mixture using inverse temperature levels β = {1, 0.1, 0.05, 0.005} respectively
+3

Références

Documents relatifs

Elle définit les principes généraux communs qui sont au nombre de cinq (par.. La médiation est un procédé librement consenti et auteur comme victime doivent pouvoir revenir sur

We do not use this intrusive type of stochastic method in this paper, in- stead, we advocate a compromise between non-intrusive and intrusive methods; we consider the

Les variables locales sont placées sur la pile en début de procédure et libérées en fin. Exemple d ’utilisation de

The performance of the methods was compared in terms of their Monte Carlo variances and in terms of length of computing time to achieve a given Monte Carlo variance. The

[r]

Abstract.- Employing the Monte Carlo technique hopping at the Fermi level in a cubic lattice has been simulated as a function of temperature, electric field, hopping parameter 2a,

Recently a method has been developed for getting to very high moments (up to the 12th and beyond) in cases where a &#34;collective&#34; state-vector (i.e. a state-vector

Simulate one game until a node (=position) L of T (thanks to bandit algorithms); put this simulation S in memory..